You are on page 1of 20

Chapter 2

Rolling Motion; Angular Momentum

2.1 The Important Stuff


2.1.1 Rolling Without Slipping
When a round, symmetric rigid body (like a uniform cylinder or sphere) of radius R rolls
without slipping on a horizontal surface, the distance though which its center travels (when
the wheel turns by an angle θ) is the same as the arc length through which a point on the
edge moves:
∆xCM = s = Rθ (2.1)

These quantities are illustrated in Fig. 2.1.


dxCM
The speed of the center of mass of the rolling object, vCM = dt
and its angular speed
are related by
vCM = Rω (2.2)

and the acceleration magnitude of the center of mass is related to the angular acceleration
by:
aCM = Rα (2.3)

s
DxCM q
R
s=Rq

s
Figure 2.1: Illustration of the relation between ∆x, s, R and θ for a rolling object.

35
36 CHAPTER 2. ROLLING MOTION; ANGULAR MOMENTUM

The kinetic energy of the object is:

Kroll = 12 ICM ω 2 + 12 MvCM


2
. (2.4)

The first term on the right side represents the rotational kinetic energy of the object about
its symmetry axis; the second term represents the kinetic energy the object would have if
it moved along with speed vCM without rotating (i.e. just translational motion). We can
remember this relation simply as: Kroll = Krot + Ktrans .
When a wheel rolls without slipping there may be a frictional force of the surface on
the wheel. If so, it is a force of static friction (which does no work) and depending on the
situation it could point in the same direction or opposite the motion of the center of mass;
in all cases it tends to oppose the tendency of the wheel to slide.

2.1.2 Torque as a Vector (A Cross Product)


In the last chapter we gave a definition for the torque τ acting on a rigid body rotating
around a fixed axis. We now give a more general definition for “torque”; we define the
torque acting on a single particle (relative to some fixed point O) when a force acts on it.
Suppose the (instantaneous) position vector of a particle (relative to the origin O) is r
and a single force F acts on it. Then the torque τ acting on the particle is

τ = r×F (2.5)

If φ is the angle between the position vector r and the force F then the torque τ has
magnitude
τ = rF sin φ

2.1.3 Angular Momentum of a Particle and of Systems of Particles


There is yet more important quantity having to do with rotations that will be of help in
solving problems involving rotating objects; just as the linear momentum p was of importance
in problems with interacting particles, the angular momentum of objects which have motion
about a given axis will be useful when these objects interact with one another. Admittedly,
some of the first definitions and theorems will be rather abstract! But we will soon apply
the ideas to simple objects which rotate around an axis and then the theorems and examples
will be quite down–to–earth.
We start with a fundamental definition; if a particle has position vector r and linear
momentum p, both relative to some origin O, then the angular momentum of that particle
(relative to the origin) is defined by:

` = r × p = m(r × v) (2.6)
kg·m2
Angular momentum has units of s .
One can show that the net torque on a particle is equal to the time derivative of its
angular momentum:
2.1. THE IMPORTANT STUFF 37

X d``
τ = (2.7)
dt
This relation is analogous to the relation F = dp from linear motion.
P
dt
For a set of mass points in motion, we define the total angular momentum as the (vector)
sum of the individual angular momenta:

L = ` 1 + `2 + ` 3 + . . .
When we consider the total angular momentum, we can prove a theorem which is a bit
different in its content than Eq. 2.7. It’s a bit subtle; when the particles in a system all move
around they will be acted upon by forces from outside the system but also by the forces
they all exert on one another. What the theorem says is that the rate of change of the total
angular momentum just comes from the torques arising from forces exerted from outside the
system. This is useful because the external torques aren’t so hard to calculate.
The theorem is:
X dL
τ ext = (2.8)
dt
This tells us that when the sum of external torques is zero then L is constant (conserved).
We will encounter this theorem most often in problems where there is rotation about a fixed
axis (and then once again we will only deal with the z components of τ and L).

2.1.4 Angular Momentum for Rotation About a Fixed Axis


An extended object is really a set of mass points, and it has a total angular momentum
(vector) about a given origin. We will keep things simple by considering only rotations
about an axis which is fixed in direction (say, the z direction), and for that case we only
need to consider the component of L which lies along this axis, Lz . So, for rotation about a
fixed axis the “angular momentum” of the rigid object is (for our purposes) just a number ,
L. Furthermore, one can show that if the angular velocity of the object is ω and its moment
of inertia about the given axis is I, then its angular momentum about the axis is

L = Iω (2.9)
Again, there is a correspondence with the equations for linear motion:
px = mvx ⇔ L = Iω

2.1.5 The Conservation of Angular Momentum


In the chapter on Momentum (in Vol. 1) we used an important fact about systems for which
there is no (net) external force acting: The total momentum remains the same. One can
show a similar theorem which concerns net external torques and angular momenta.
For a system on which there is no net external torque, the total angular momentum
remains constant: Li = Lf . This principle is known as the Conservation of Angular
Momentum.
38 CHAPTER 2. ROLLING MOTION; ANGULAR MOMENTUM

2.2 Worked Examples


2.2.1 Rolling Without Slipping

1. An automobile traveling 80.0 km/hr has tires of 75.0 cm diameter. (a) What
is the angular speed of the tires about the axle? (b) If the car is brought to a
stop uniformly in 30.0 turns of the tires (without skidding), what is the angular
acceleration of the wheels? (c) How far does the car move during the braking?
[HRW5 12-3]

(a) We know that the speed of the center of mass of each wheel is 80.0 km/hr. And the
radius of each wheel is R = (75.0 cm)/2 = 37.5 cm. Converting the speed to ms , we have:

103 m
! !
km

km
 1h
80 h
= 80 h
= 22.2 ms
3600 s 1 km

From the relation between vCM and ω for an object which rolls without slipping, we have:
vCM
vCM = ωR =⇒ ω=
R
and we get
(22.2 ms )
ω= = 59.3 rad
s
(0.375 m)
The angular speed of the wheel is 59.3 rad
s
.
(b) As the car comes to a halt, the tires go through 30.0 turns. Thus they have an angular
displacement of (with θ0 = 0):
!
2π rad
θ = (30.0 rev) = 188.5 rad .
1 rev

Also, when the wheel has come to a halt, its angular velocity is zero!
So we have the initial and final angular velocities and the angular displacement. We can
get the angular acceleration of the wheel from Eq. 1.8. From that equation we get:

ω 2 − ω02 (0 rad )2 − (59.3 rad )2


α= = s s
= −9.33 rad
s2
2θ 2(188.5 rad)

The magnitude of the wheels’ angular acceleration is 9.33 rad s2


. The minus sign in our result
indicates that α goes in the sense opposite to that of the initial angular velocity (and angular
displacement) of the wheel during the stopping.
2.2. WORKED EXAMPLES 39

10 N N 10 N

fs
Mg

(a) (b)
Figure 2.2: (a) Constant horizontal applied to a rolling wheel in Example 3. (b) The forces acting on the
wheel, with the points of application as indicated.

(c) As we saw, the angular displacement of any wheel during the stopping was 188.5 rad.
The radius of the wheel is R = 0.375 m, so from Eq. 2.1 the linear displacement of the wheel
(i.e. its center) is:
xCM = Rθ = (0.375 m)(188.5 rad) = 70.7 m
so the car goes 70.7 m before coming to a halt.

2. A bowling ball has a mass of 4.0 kg, a moment of inertia of 1.6 × 10−2 kg · m2 and
a radius of 0.10 m. If it rolls down the lane without slipping at a linear speed of
4.0 ms , what is its total energy? [Ser4 11-5]

The total (kinetic) energy of an object which rolls without slipping is given by Eq. 2.4.
To use this equation we have everything we need except the angular speed of the ball. From
Eq. 2.2 it is related to the linear velocity of the ball by vCM = Rω, so the angular speed is

vCM (4.0 ms )
ω= = = 40.0 rad
s
R (0.10 m)

and then the kinetic energy is

Kroll = 21 ICM ω 2 + 12 MvCM


2

= 12 (1.6 × 10−2 kg · m2 )(40.0 rad


s
)2 + 21 (4.0 kg)(4.0 ms )2
= 44.8 J

The total kinetic energy of the ball is 44.8 J.

3. A constant horizontal force of 10 N is applied to a wheel of mass 10 kg and


radius 0.30 m as shown in Fig. 2.2. The wheel rolls without slipping on the
horizontal surface, and the acceleration of its center of mass is 0.60 sm2 . (a) What
are the magnitude and direction of the frictional force on the wheel? (b) What
is the rotational inertia of the wheel about an axis through its center of mass
and perpendicular to the plane of the wheel? [HRW5 12-9]
40 CHAPTER 2. ROLLING MOTION; ANGULAR MOMENTUM

(a) The forces which act on the wheel along with where these forces are applied are shown
in Fig 2.2 (b). In addition to the applied force of 10 N which points to the right, there is
a force of static friction between the surface and the wheel (of magnitude fs ), which for
now we draw pointing to the left (we can ask: Does it really point that way?). There are
vertical forces acting on the wheel (from gravity and the normal force of the surface) but
these clearly cancel out and for now we don’t need to worry about them.
Even though the wheel will be rolling during its motion, Newton’s 2nd law still holds,
and the sum of the horizontal forces gives max. Here the wheel is clearly accelerating to the
right and so with the choice of directions given in the figure, we find:

Fx = 10.0 N − fs = max = (10 kg)(0.60 sm2 ) = 6.0 N


X

so that
fs = 10.0 N − 6.0 N = 4.0 N
and since this is positive, the frictional force does indeed point to the left, as we guessed.
Actually, it wasn’t so hard to guess that, since only a leftward frictional force could make
the wheel rotate clockwise —as we know it must here— but for some problems in rolling
motion, the direction of the static friction force may not be so evident.
(b) Rotational inertia is related to net torque and angular acceleration by way of τ = Iα.
It is true that in this problem the rotating object is also accelerating but it turns out this
relation still holds as long as we choose the center of mass to be the rotation axis.
Considering the four forces in Fig. 2.2 (b), the applied force and the (effective) force of
gravity are applied at the center of the wheel, so they give no torque about its center. The
normal force of the surface is applied at the rim, but its direction is parallel to the line which
joins the application point to the center so it too gives no torque. All that remains is the
friction force, applied at a distance R from the center and perpendicular to the line joining
this point and the axis; this force gives a clockwise rotation, so if we take the clockwise
direction as the positive sense for rotations, then the net torque on the wheel is

τ = +fs r = (4.0 N)(0.30 m) = 1.2 N · m

From Eq. 2.3 we know the angular acceleration of the wheel; it is


aCM (0.60 sm2 )
α= = = 2.0 rad
s2
.
R (0.30 m)
and then from τ = Iα we get the rotational inertia:
τ (1.2 N · m)
I= = = 0.60 kg · m2
α (2.0 rad
s 2 )

4. A round, symmetrical object of mass M, radius R and moment of inertia I


rolls without slipping down a ramp inclined at an angle θ. Find the acceleration
of its center of mass.
2.2. WORKED EXAMPLES 41

I, M, R N

fs Mg sin q
q

Mg cos q

Mg

Figure 2.3: Round, symmetrical object with mass M , radius R and moment of inertia I rolls down a ramp
sloped at angle θ from the horizontal.

The problem is diagrammed in Fig. 2.3. We show the forces acting on the object and
where they are applied. The force of gravity, Mg is (effectively) applied at the center
of the object. As usual we decompose this force into its components down the slope and
perpendicular to the slope. The slope exerts a normal force N at the point of contact. Finally
there is a force of static friction fs from the surface; this force points along the surface and
we can pretty quickly see that it must point up the slope because it is the friction force
which gives the object an angular acceleration, which (here) is in the clockwise sense.
Apply Newton’s 2nd law first: The forces perpendicular to the slope cancel, so that
N = Mg cos θ (but we won’t need this fact). If aCM is the acceleration of the CM of the
object down the slope, then adding the forces down the slope gives

Mg sin θ − fs = MaCM (2.10)

Now we look at the net torque on the rolling object about its center of mass. The force
of gravity acts at the center, so it gives no torque. The normal force of the surface acts at
the point of contact, but since it is parallel to the line joining the pivot and the point of
application, it also gives no torque. The force of friction is applied at a distance R from the
pivot and it is perpendicular to the line joining the pivot and point of application. So the
friction force gives a torque of magnitude

τ = Rfs sin 90◦ = Rfs .

and if we take the clockwise sense to be positive for rotations, then the net torque on the
object about its CM is
τnet = Rfs
From the relation τ = Iα we then have

τ = Rfs = Iα (2.11)

but we can also use the fact that for rolling motion (without slipping) the linear acceleration
of the CM and the angular acceleration are related by:
aCM
aCM = rα =⇒ α=
R
42 CHAPTER 2. ROLLING MOTION; ANGULAR MOMENTUM

and using this in Eq. 2.11 gives


IaCM IaCM
Rfs = =⇒ fs = (2.12)
R R2
where we choose to isolate fs (the magnitude of the friction force) so that we can put the
result into Eq. 2.10. When we do that, we get:
IaCM
Mg sin θ − = MaCM
R2
Now solve for aCM :
IaCM I
 
Mg sin θ = MaCM + 2
= M + 2 aCM
R R
Mg sin θ
aCM =   .
M + RI2
If we divide top and bottom of the right side by M, this can be written:
g sin θ
aCM =   .
1 + MIR2

Our result is sensible in that if I is very small then aCM is nearly equal to g sin θ, the
result for a mass sliding with no rolling motion.

5. A uniform sphere rolls down an incline. (a) What must be the incline angle
if the linear acceleration of the center of the sphere is to be 0.10 g? (b) For this
angle, what would be the acceleration of a frictionless block sliding down the
incline? [HRW5 12-7]

(a) We will use the formula for aCM (for rolling without slipping down a slope) in a previous
problem. Note that we are not given the mass of the sphere! But it turns out that we don’t
need it, because for a uniform sphere, we have
I 2
2
=
MR 5
and as we can see from our earlier result,
g sin θ
aCM =   ,
1 + MIR2

aCM just depends on the combination I/(MR2 ). Solving this equation for sin θ and plugging
in the given numbers, we get:
aCM I (0.10g)
 
sin θ = 1+ = (1 + 25 ) = (0.10) 75 = 0.14
g MR2 g
2.2. WORKED EXAMPLES 43

Which gives us
θ = sin−1 (0.14) = 8.0◦

(b) As we saw in the Chapter on forces (Volume 1) when a mass slides down a frictionless
incline its linear acceleration is given by a = g sin θ. For the slope angle found in part (a),
this is
a = g sin θ = (9.80 sm2 ) sin 8.0◦ = 1.4 sm2
We can also calculate a/g:
a
= sin θ = sin 8.0◦ = 0.14
g
so for the frictionless case we have a = 0.14g.

2.2.2 Torque as a Vector (A Cross Product)

6. What are the magnitude and direction of the torque about the origin on
a plum (!) located at coordinates (−2.0 m, 0, 4.0 m) due to force F whose only
component is (a) Fx = 6.0 N, (b) Fx = −6.0 N, (c) Fz = 6.0 N, and (d) Fz = −6.0 N ?
[HRW5 12-21]

(a) The (vector) torque on a point particle is given by


τ = r×F
I find it easiest to set up the cross product in determinant notation, discussed in Chapter
1 of Volume 1. We note that the units of the result must be N · m; then the cross product
of r and F is

i j k

r×F= −2.0 0.0 4.0 N · m = (+(4.0)(6.0) N · m)j = (+24.0 N · m)j

6.0 0.0 0.0

The torque τ has magnitude 24.0 N · m and points in the +y direction.


(b) Is is fairly clear that if we had had Fx = −6.0 N in part (a), we would have gotten
τ = (−24.0 N · m)j
so the torque would have magnitude 24.0 N · m and point in the −y direction.
(c) When the only component of F is Fz = 6.0 N, then we have

i j k

r×F = −2.0 0.0 4.0 N · m = (−(−2.0)(6.0) N · m)j = (+12.0 N · m)j

0.0 0.0 6.0

so the torque would have magnitude 12.0 N · m and point in the +y direction.
(d) If instead we have only Fz = −6.0 N then the sign of the result in part (c) changes, and
the torque would have magnitude 24.0 N · m and point in the −y direction.
44 CHAPTER 2. ROLLING MOTION; ANGULAR MOMENTUM

6.5 kg
2.2 m/s

1.5 m 3.6 m/s

O 2.8 m
3.1 kg

Figure 2.4: Two masses and their motion relative to the point O, as in Example 8.

2.2.3 Angular Momentum of a Particle and of Systems of Particles

7. The position vector of a particle of mass 2.0 kg is given as a function of time by


r = (6.0i + 5.0tj) m when t is given in seconds). Determine the angular momentum
of the particle as a function of time. [Ser4 11-17]

The angular momentum of a point mass is given by ` = r×p. The velocity of our particle
is given by
d d
v = r = (6.0i + 5.0tj) m = (5.0j) ms
dt dt
and its momentum is

p = mv = (2.0 kg)(5.0j) ms = (10.0j) kg·m


s

Now take the cross product to get ` :



i j k

kg·m2 2
` = r×p = 6.0 5.0t 0.0 = (60.0k) kg·m

s s
0.0 10.0 0.0

8. Two objects are moving as shown in Fig. 2.4. What is their total angular
momentum about point O? [HRW5 12-27]

In this problem we must use the definition of the angular momentum of a particle (with
respect to some origin O):
` = r×p
First consider the 3.1 kg mass. Its momentum vector has magnitude

p1 = m1 v1 = (3.1 kg)(3.6 ms ) = 11.2 kg·m


s
2.2. WORKED EXAMPLES 45

and is directed upward in this picture. The vector r which goes from point O to the particle
has magnitude 2.8 m and points to the right. By the right–hand rule for cross products, the
vector r × p points up out of the page, which is along the +z axis; and since the two vectors
are perpendicular, the magnitude of r × p is
2
|r1 × p1 | = r1 p1 sin 90◦ = (2.8 m)(11.2 kg·m
s
) = 31.2 kg·m
s

and so the angular momentum of this particle about O is


2
` 1 = (+31.2 kg·m
s
)k

The 6.5 kg mass has a momentum vector of magnitude

p2 = m2 v2 = (6.5 kg)(2.2 ms ) = 14.3 kg·m


s

and is directed to the right. The vector r which goes from the point O to the particle has
magnitude 1.5 m and points straight up. By the right–hand rule for cross products, the
vector r × p points into the page, which is along the −z axis; and since the two vectors are
perpendicular, the magnitude of r × p is
2
|r2 × p2 | = r2 p2 sin 90◦ = (1.5 m)(14.3 kg·m
s
) = 21.5 kg·m
s

and so the angular momentum of this particle about O is


2
`2 = (−21.5 kg·m
s
)k

The total angular momentum of the system is


2 2
` = `1 + ` 2 = (31.2 − 21.5)k kg·m
s
= (+9.8 kg·m
s
)k

2.2.4 Angular Momentum for Rotation About a Fixed Axis

9. A uniform rod rotates in a horizontal plane about a vertical axis through one
end. The rod is 6.00 m long, weighs 10.0 N, and rotates at 240 rev/min clockwise
when seen from above. Calculate (a) the rotational inertia of the rod about the
axis of rotation and (b) the angular momentum of the rod about that axis. [HRW5
12-45]

(a) The mass of the rod is

M = W/g = (10.0 N)/(9.80 sm2 ) = 10.2 kg


rad
and its angular velocity in units of s
is
!
2π rad 1 min

rev
ω = 240 min
= 25.1 rad
s
1 rev 60 s
46 CHAPTER 2. ROLLING MOTION; ANGULAR MOMENTUM

We know the formula for the moment of inertia of a uniform rod rotating about an axis at
one of its ends (see Chapter 1, Fig. 1) so we calculate I as:

I = 13 ML2 = 13 (10.2 kg)(6.00 m)2 = 12.2 kg · m2

The rotational inertia of the rod (about the given axis) is 12.2 kg · m2.
(b) The angular momentum of the rotating rod will be given by L = Iω. We find that the
magnitude of the angular momentum is:
2
L = Iω = (12.2 kg · m2)(25.1 rad
s
) = 309 kg·m
s

The vector L would point upward (along the +z axis if the rotation were counterclockwise
as seen from above. That is not the case (it is clockwise) so the direction of L is downward.

2.2.5 The Conservation of Angular Momentum

10. Suppose that the Sun runs out of nuclear fuel and suddenly collapses to form
a white dwarf star, with a diameter equal to that of the Earth. Assuming no
mass loss, what would then be the Sun’s new rotation period, which currently
is about 25 days? Assume that the Sun and the white dwarf are uniform, solid
spheres; the present radius of the Sun is 6.96 × 108 m. [HRW5 12-55]

In this simplified account of what will happen to our Sun, its radius will decrease without
any interactions from other masses, including torques. Without external torques, the angular
momentum of the Sun will remain the same, even if it suddenly shrinks to a much smaller
size.
The present angular velocity of the Sun (using the given data) is
! ! !
1 rev 2π rad 1 day rad
ωi = = 2.91 × 10−6 s
25 day 1 rev 86400 s

and assuming it is a uniform sphere (a bad assumption, actually) its present moment of
inertia is
Ii = 25 MR2i
(we’ll leave it in this form; later on, the value of M will cancel out). Its initial angular
momentum is
Li = Ii ωi = 25 MR2i (2.91 × 10−6 rad
s
) .
After the Sun shrinks, it has a new (much smaller) radius, but the same mass, M. Its
new moment of inertia is
If = 25 MR2f
and if its final angular velocity is ωf , then its final angular momentum is

Lf = If ωf = 25 MR2f ωf
2.2. WORKED EXAMPLES 47

I= 250 kg m2

2.0 m

10 rev/min w=?

25 kg

(a) (b)

Figure 2.5: Child jumps onto a rotating merry-go-round in Example 11.

Conservation of angular momentum gives us Li = Lf and so:


2
5
MR2i (2.91 × 10−6 rad
s
) = 25 MR2f ωf .

We cancel lotsa things and find:


R2i rad
ωf = (2.91 × 10−6 )
R2f s

The present radius of the Sun is 6.96 × 108 m and the radius of the Earth is 6.37 × 106 m so
we find that the angular velocity of the Sun will be

(6.96 × 108 m)2 rad rad


ωf = 6 2
(2.91 × 10−6 s
) = 3.47 × 10−2 s
.
(6.37 × 10 m)
To get the period of the Sun’s motion, use
ω 1 2π
f= =⇒ T = =
2π f ω
So:
2π 2π 1 min
 
Tf = = rad
= 181 s = 3.0 min
ωf (3.47 × 10−2 s
) 60 s
The squooshed–down Sun will have a rotation period of 3.0 minutes!

11. A merry-go-round of radius R = 2.0 m has a moment of inertia I = 250 kg · m2


and is rotating at 10 rev/min. A 25 kg child jumps onto the edge of the merry-go-
round. What is the new angular speed of the merry-go-round? [Ser4 11-28]

We diagram the problem in Fig. 2.5. In (a) the child is waiting (motionless) by the rim
of the rotating wheel and in (b) he/she/it has just stepped on. What do we know about
pictures (a) and (b)? We know that if we consider the “system” to be the combination of
merry-go-round and child, if the child just steps onto the wheel at its rim there will be no
48 CHAPTER 2. ROLLING MOTION; ANGULAR MOMENTUM

external torques on this system. And so the total angular momentum of the system will be
conserved.
We calculate the total angular momentum that we see in Fig. 2.5 (a). Only the wheel is
in motion; its angular velocity is
!
2π rad 1 min

rev
ω = (10 min ) = 1.05 rad
s
1 rev 60 s

so its angular momentum is


2
L = Iω = (250 kg · m2)(1.05 rad
s
) = 262 kg·m
s

The child is motionless, so this is the total initial angular momentum of the system: Li =
2
262 kg·m
s
.
After the child steps onto the merry-go-round (and stays at the same place near its outer
edge) we have a rotating system with a different moment of inertia. The child is (basically)
a point mass at a distance of 2.0 m from the rotation axis, so the new moment of inertia is
found from summing the moments of the original wheel and the child:

I 0 = (250 kg · m2) + MchildR2 = (250 kg · m2) + (25.0 kg)(2.0 m)2 = 350 kg · m2

and if ω 0 is the final angular velocity, then the final angular momentum is given by:

Lf = I 0ω 0

But from angular momentum conservation, Li = Lf and this lets us solve for ω 0 :
2
Li = 262 kg·m
s
= Lf = I 0ω 0 = (350 kg · m2)ω 0

which gives
2
(262 kg·m )
0
ω = s
= 0.748 rad
s
(350 kg · m2)
We can convert this back to revolutions per minute to give:
1 rev 60 s
  
ω 0 = (0.748 rad
s
) rev
= 7.14 min .
2π rad 1 min
rev
The merry-go-round slows down to 7.14 min after the child steps on.

12. Two astronauts (see Fig. 2.6) each having a mass of 75 kg are connected by a
10 m rope of negligible mass. They are isolated in space, orbiting their center of
mass at speeds of 5.0 ms . Calculate (a) the magnitude of the angular momentum of
the system by treating the astronauts as particles and (b) the rotational energy
of the system. By pulling on the rope, the astronauts shorten the distance
between them to 5.0 m. (c) What is the new angular momentum of the system?
(d) What are their new speeds? (e) What is the new rotational energy of the
2.2. WORKED EXAMPLES 49

CM

10 m

Figure 2.6: Two astronauts out in space, playing around with a rope, in Example 12.

system? (f) How much work is done by the astronauts in shortening the rope?
[Ser4 11-53]

(a) Each astronaut in Fig. 2.6 has a mass of 75 kg, a speed of 5.0 ms and moves perpendicularly
to the line which joins the origin (i.e. the CM) to their locations. Each one is 5.0 m from
the origin, so the magnitude of the angular momentum of each astronaut is
kg·m2
` = rp sin 90◦ = (5.0 m)(75 kg)(5.0 ms ) = 1.88 × 103 s

(By the right–hand–rule, the angular momentum vector for each one points out of the page.
Since the rope is taken to be massless, the sum of the angular momenta for the astronauts
is the L for the system; these just add together to give:
kg·m2 kg·m2
L = 2(1.88 × 103 s
) = 3.75 × 103 s

(b) The rotational energy of the system is simply the total kinetic energy of the astronauts
(again, the rope is massless). This is:
 
K =2 1
2
(75 kg)(5.0 ms )2 = 1.88 × 103 J = 1.88 kJ

(c) As the astronauts pull on the rope to decrease their separation, there are all kinds of
internal forces in the astronaut–rope system, but there are no external torques on the system.
As a result, the total angular momentum stays the same: Its magnitude is still
kg·m2
L = 3.75 × 103 s

(d) In the new configuration, each astronaut will be 5.0 m/2 = 2.5 m from the center of
rotation and will still have a velocity (and linear momentum) perpendicular to the line joining
the rotation center to his location. If the new linear momentum of each astronaut is now p0 ,
then we can use the expression for the total angular momentum in the new configuration to
write:
2
L0 = L = 3.75 × 103 kg·ms
= 2(r0 p0 ) = 2(2.5 m)p0
50 CHAPTER 2. ROLLING MOTION; ANGULAR MOMENTUM

q m
M, l

Figure 2.7: Odd contraption where a sliding particle sticks to the end of a rod which is pivoted about O,
in Example 13.

and then solve for p0 :


2
(3.75 × 103 kg·m )
0
p = s
= 750 kg·m
s
2(2.5 m)
and from this we can get the new speed of each astronaut:

p0 (750 kg·m )
0
v = = s
= 10.0 ms
m (75.0 kg)

(e) The total kinetic energy of the astronauts is now


 
K0 = 2 1
2
(75.0 kg)(10.0 ms )2 = 7.50 × 103 J = 7.50 kJ

(f) Our answer for (e) is larger (in fact, quite a bit larger!) than the initial kinetic energy
which we calculated in part (b). The difference had to come from the work done by astronauts
in pulling on the rope. So the work done was

W = K 0 − K = 7.50 kJ − 1.88 kJ = 5.62 kJ .

The astronauts did 5.62 kJ of work in shortening the length of rope between them.

13. The particle of m in Fig. 2.7 slides down the frictionless surface and collides
with the uniform vertical rod, sticking to it. The rod pivots about O through
the angle θ before momentarily coming to rest. Find θ in terms of the other
parameters given in the figure. [HRW5 12-69]

To make any sense of the motion of the small mass and the rod we need to go through
this collision step–by–step to see which of physical principles can be applied.
The mass starts at a height h above the level part of the surface and slides down the
smooth slope. Just before it hits the rod, we know that the mechanical energy of the system
has been conserved because no friction–type force have been acting. This will allow us to
find the speed of the mass at the bottom of the ramp.
2.2. WORKED EXAMPLES 51

w
l
v

(a) (b)

Figure 2.8: (a) Just before mass strikes end of rod it has linear motion, with speed v. (b) After mass sticks
to rod, both rod and mass rotate about the pivot with angular velocity ω.

Next, the mass has a very brief (and very sticky) encounter with the end of the rod. We
are used to treating a set of masses as being isolated during the short time that they exert
forces on one another. And indeed during the collision the mass and rod are isolated, in a
sense. But it is not true that we can ignore the external forces during their brief encounter,
because in fact the pivot will exert a significant force on the rod that can’t be ignored. But
it is true that we can ignore the external torques that act on the rod–mass system (about
the pivot, that is) because the pivot force will give no torque. And because of this we will be
able to use the fact that the total angular momentum is conserved during the rapid, sticky
collision of the mass and the end of the rod.
Finally, after the mass is stuck to the end of the rod, the rod makes a swing upward,
momentarily coming to rest. During this part of the motion it is total mechanical energy
which is conserved – as long as there is no friction in the pivot!
Now we start writing down some equations to express the principles laid out here.
First we apply the condition of energy conservation between the initial position of the
mass and the instant before it hit the rod. At first, the mass has only gravitational energy:
Ei = mgh. Just before the collision, it has only kinetic energy; if its speed there is v, then
Ef = 21 mv 2 and energy conservation, Ei = Ef gives us:
q
mgh = 21 mv 2 =⇒ v= 2gh (2.13)

Now we look at the rapid collision, in which angular momentum is conserved. Just before
the mass m hits the end of the rod, its motion is horizontal (perpendicular to line joining the
pivot and its location) and it has the speed v which we have already calculated, as shown in
Fig. 2.8) (a). At this time its distance from the pivot is l, so its angular momentum about
the pivot (in the clockwise sense) is Li = lmv. Since at this time the rod is motionless, that
is the total angular momentum of the mass–rod system.
After the collision the mass and rod form a rotating system with moment of inertia
M
 
I= 1
3
Ml2 2
+ ml = + m l2
3
since the mass m is assumed small and is fixed at a distance l from the pivot. If the angular
52 CHAPTER 2. ROLLING MOTION; ANGULAR MOMENTUM

l q
l cos q
CM

Figure 2.9: Some geometry for the “final” position of the mass/rod combination.

velocity of the rod/mass just after the collision is ω, then the final angular momentum is

M
 
Lf = Iω = + m l2ω
3
Conservation of angular momentum, Li = Lf gives us:

M
 
lmv = + m l2ω (2.14)
3
Finally, after the collision we have energy conservation all during the upward swing of
the rod/mass. (We will now let the letters i and f refer to initial and final positions for
this part.) Just after the mass has stuck to the rod, the kinetic energy of the system is its
rotational kinetic energy,
M
 
Ki = 12 Iω 2 = 21 + m l2 ω2 .
3
To calculate potential energies we will measure height from the “ground” level. The mass
is small and basically has no height but the CM of the rod is at a height 2l so the initial
potential energy is
l
Ui = Mg
2
At the top of the swing, we know what the final kinetic energy is: It’s zero! As for the
final potential energy we will need to do a little geometry, as in Fig. 2.9. From this figure
we can see that the height of the small mass is now l − l cos θ. The vertical distance of the
rod’s CM downward from the pivot is 2l cos θ, so that its height is now l − 2l cos θ. Then the
final potential energy is (after a little factoring):

Uf = mgl(1 − cos θ) + Mgl(1 − 12 cos θ)

The change in potential energy is


l
∆U = Uf − Ui = mgl(1 − cos θ) + Mgl(1 − 12 cos θ) − Mg
2
2.2. WORKED EXAMPLES 53

= mgl(1 − cos θ) + Mgl( 12 − 12 cos θ)


Mgl
= mgl(1 − cos θ) + (1 − cos θ)
2
M
 
= m+ gl(1 − cos θ)
2
and the change in kinetic energy is just
M
 
∆K = Kf − Ki = − 12 + m l2ω2
3
so conservation of energy, ∆K + ∆U = 0 gives us:
M M
   
m+ gl(1 − cos θ) − 12 + m l2ω2 = 0 (2.15)
2 3
Have we done enough physics to get us to the answer? In the above equations, the
unknowns are v, ω and θ, and we do have three equations: 2.13, 2.14 and 2.15. So we can
solve for them; in particular we can find θ which is what the problem asks for. (The answer
will be expressed in terms of M, m, and l which we take as given.)
Here’s one way we can solve them. In Eq. 2.15 we will need ω 2 , so we can write 2.14 as:
mv
ω = M 
3
+ m l
then square it and use v 2 = 2gh from 2.13:
2 m2 v 2 2m2 gh
ω = 2 = 2
M M
3
+m l2 3
+m l2
Using this, we can substitute for ω 2 in Eq. 2.15 to get
M M 2m2 gh
   
m+ gl(1 − cos θ) = 1
2
+ m l2  2
2 3 M
+ m l2 3

There is much to cancel here! Some algebra gives us:


m2 h
(1 − cos θ) =  M

M

l m+ 2 3
+m
Don’t despair; we’re nearly home! Multiplying the top and bottom of the right hand side by
6 tidies things up to give:
6m2 h
(1 − cos θ) =
l(2m + M)(M + 3m)
Isolate cos θ to get:
6m2 h
cos θ = 1 −
l(2m + M)(3m + M)
And finally:
6m2 h
" #
−1
θ = cos 1−
l(2m + M)(3m + M)
Finished!
54 CHAPTER 2. ROLLING MOTION; ANGULAR MOMENTUM

You might also like